Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

JTTEE5 20:817–828

DOI: 10.1007/s11666-011-9632-2
1059-9630/$19.00  ASM International

Peer Reviewed
Column Formation in Suspension
Plasma-Sprayed Coatings and Resultant
Thermal Properties
Kent VanEvery, Matthew J.M. Krane, Rodney W. Trice, Hsin Wang, Wallace Porter, Matthew Besser,
Daniel Sordelet, Jan Ilavsky, and Jonathan Almer

(Submitted November 15, 2010; in revised form February 15, 2011)

The suspension plasma spray (SPS) process was used to produce coatings from yttria-stabilized zirconia
(YSZ) powders with median diameters of 15 lm and 80 nm. The powder-ethanol suspensions made with
15-lm diameter YSZ particles formed coatings with microstructures typical of the air plasma spray
(APS) process, while suspensions made with 80-nm diameter YSZ powder yielded a coarse columnar
microstructure not observed in APS coatings. To explain the formation mechanisms of these different
microstructures, a hypothesis is presented which relates the dependence of YSZ droplet flight paths on
droplet diameter to variations in deposition behavior. The thermal conductivity (kth) of columnar SPS
coatings was measured as a function of temperature in the as-sprayed condition and after a 50 h, 1200 °C
heat treatment. Coatings produced from suspensions containing 80 nm YSZ particles at powder con-
centrations of 2, 8, and 11 wt.% exhibited significantly different kth values. These differences are con-
nected to microstructural variations between the SPS coatings produced by the three suspension
formulations. Heat treatment increased the kth of the coatings generated from suspensions containing 2
and 11 wt.% of 80 nm YSZ powder, but this kth increase was less than has been observed in APS
coatings.

Keywords APS coatings, PS microstructures, suspension An EB-PVD YSZ coating consists of a collection of
plasma spray, thermal properties columnar, quasi-single crystals (Ref 6, 7). Typical depo-
sition conditions result in these crystals growing primarily
normal to the substrate surface, which is located directly
above the flux emitted by the vaporized feedstock (Ref 8,
9). The resulting columnar grains, which tend to extend
1. Introduction from the substrate to the coating top surface, usually have
diameters on the order of 10 lm (Ref 6, 7). These grains
Yttria-stabilized zirconia (YSZ) coatings are widely contain closed and open pores and are separated from
used in gas turbines to insulate metallic components from neighboring columns by a continuous gap that varies in
the heat generated during engine operation (Ref 1). These width from a few nanometers at the substrate to ~1 lm
thermal barrier coatings (TBCs) are commonly applied near the coating top surface (Ref 6, 7).
through either the electron beam physical vapor deposi- An APS coating is deposited by liquid droplets that
tion (EB-PVD) or air plasma spray (APS) processes impact a substrate, spread across the surface, and solidify
(Ref 1, 2). In the case of EB-PVD, electron beam heating into disc-like structures referred to as lamellae (Ref 3, 4,
of a YSZ ingot produces a vapor that subsequently con- 10). During droplet spreading the liquid may not com-
denses on the substrate (Ref 1, 2). In APS, YSZ powder is pletely wet the underlying surface, which will result in
injected into a DC plasma which melts the powder parti- interlamellar pores (Ref 3, 4). Despite the presence of
cles while propelling them onto the substrate (Ref 1-3). these interlamellar pores, heat is transferred out of a
The differences in the two processes lead to distinct spread droplet mainly through the boundary in contact
coating microstructures (Ref 4-7). with the substrate or previously deposited coating (Ref 4,
11, 12). Consequently, as a lamella solidifies, columnar
grains grow perpendicular to this boundary (Ref 10-12).
The contraction accompanying lamella solidification and
Kent VanEvery, Progressive Surface, Grand Rapids, MI; cooling generates strains that can form cracks (Ref 4, 11,
Matthew J.M. Krane and Rodney W. Trice, School of Materials 12). Since the bonding within grains is stronger than be-
Engineering, Purdue University, West Lafayette, IN; Hsin Wang tween them, these intralamellar cracks tend to propagate
and Wallace Porter, Oak Ridge National Laboratory, Oak
Ridge, TN; Matthew Besser, Ames Laboratory, Ames, IA;
along the grain boundaries, cutting the lamella. An APS
Daniel Sordelet, Caterpillar, Peoria, IL; and Jan Ilavsky and YSZ coating is thus composed of lamellae, interlamellar
Jonathan Almer, Argonne National Laboratory, Argonne, IL. porosity mainly oriented with radial axes parallel to the
Contact e-mail: rtrice@purdue.edu. substrate, and intralamellar cracks aligned normal to the

Journal of Thermal Spray Technology Volume 20(4) June 2011—817


substrate (Ref 4, 10, 12). In the literature cited above, the submicron feedstocks, the SPS process has the potential to
Peer Reviewed

APS microstructure is frequently presented as an arche- generate coatings in which the density of interlamellar
type describing all coatings produced by using a plasma to pores is significantly larger than what is common in APS
melt and deposit YSZ powder; however, as will be pre- coatings. Such a change is likely to affect the coating
sented here, such a description may not apply to the thermal conductivity. At 1250 C, which is close to the
coatings resulting from plasma spraying YSZ powders maximum gas temperature at the inlet of a commercial
with single digit micron and smaller diameters. aircraft turbine engine, photon transport comprises 10%
In APS, the feedstock particles are introduced into the of the total heat flux through a 6-8 wt.% YSZ ceramic
plasma by a carrier gas (Ref 3). As the particle diameters (Ref 21, 22). Therefore, the majority of the heat transfer
decrease, the carrier gas flow rate must increase in order through the YSZ TBCs in aircraft engines occurs by
to inject the less massive particles into the plasma plume phonon transport in the ceramic lattice. Increasing the
(Ref 3). For most powder materials, the injection of par- density of interlamellar pores should cause these phonons
ticles with diameters <5 lm requires carrier gas velocities to follow more tortuous paths through the microstructure,
that significantly perturb the plasma flow, reducing the which would lower the coating thermal conductivity (Ref
overall deposition efficiency (Ref 3). In addition, the 23). Consequently, SPS could provide a means of
electrostatic surface forces between micron-sized and improving the insulating performance of TBCs relative to
smaller particles can dominate the gravitational body current APS thermal barriers. The current study was
forces acting on the particles, resulting in particle undertaken to evaluate this possibility by characterizing
agglomeration and an inability to feed material through the microstructure and thermal conductivity of YSZ
the powder supply line (Ref 13). In many cases, the coatings produced from nanoscale powder feedstock.
technical issues associated with spraying micron and sub-
micron diameter powders (i.e., increased carrier gas flow
and feedstock agglomeration) can be solved by spraying 2. Experimental Procedures
suspensions (Ref 13, 14). Compared to a gas carrier, the
more massive liquid carrier can generate the momentum
2.1 Coating Production
required to inject these small powders into the plasma
stream at velocities that do not negatively impact the SPS was employed to generate coatings from suspen-
coating process. Furthermore, a liquid carrier can contain sions of two different sized YSZ powders in ethanol. The
dispersants that inhibit powder agglomeration (Ref 3, 13, particle size distributions within all of the suspensions
15). These advantages have led to the development of the employed in this study were measured with a Beckman
suspension plasma spray (SPS) technique (Ref 13, 14, 16- Coulter (Fullerton, CA) LS 230 particle size analyzer
19). (Ref 24). A density of 6.0 g/cm3 and complex refractive
In SPS, the suspension is typically injected into the index of 2.16-0.01i were used for powder size calculations
plasma plume through a nozzle that produces a coherent (Ref 24-27).
suspension stream or a stream of suspension droplets (Ref Micron feedstock suspensions were produced with a
17, 19). Aerodynamic drag forces from the plasma flow ZrO2-7 wt.% Y2O3 fused and crushed powder from H.C.
cause the initial stream or droplets to fragment into Starck, Inc (Euclid, OH) as described in Table 1. No ball
smaller, stable droplet sizes (Ref 15, 17). The droplet size milling was performed on this type of suspension. The
is decreased further as the solvent evaporates (Ref 13, 15, coatings produced from the 15-lm diameter YSZ sus-
17). To reduce the enthalpy required for vaporization, a pensions will subsequently be referred to as 8 M coatings,
volatile liquid like ethanol is commonly used for the sus- where the number represents the powder loading in wt.%
pension solvent (Ref 13, 15). Under optimal deposition and the letter represents the powder size, in this case,
conditions, the solvent in the droplet fully evaporates, micron-sized powders.
leaving individual powder particles or agglomerates that The nanofeedstock suspensions contained ZrO2-
melt prior to impacting on the substrate (Ref 13, 20). 8 wt.% Y2O3 powder synthesized by Inframat Advanced
Achieving these conditions with the micron and submi- Materials (Manchester, CT). Nanoparticle suspensions
cron droplets generated in SPS requires the use of shorter were formulated as detailed in Table 1; the resulting
standoff distances between the plasma torch exit and the
substrate than is typical for APS (Ref 20). This reduction
is necessary to decrease the occurrence of droplet solidi-
fication prior to deposition. A shorter standoff distance Table 1 Suspension formulation details
results in the substrate experiencing a greater heat flux. As
a result, SPS usually involves more substrate cooling than Powder
Suspension conc., wt.% Solventa YSZ particulate d50
is used in APS to prevent thermal stress generation within
a coating from causing spallation (Ref 13). 8M 8 Ethanol 15 ± 6 lm (volume-based)
The microstructure of plasma-sprayed coatings can 2N 2 Ethanol 85 ± 13 nm (number-based)
8N 8 Ethanol 85 ± 13 nm (number-based)
change as the feedstock powder size is reduced. For 11 N 11 Ethanol 89 ± 13 nm (number-based)
example, decreasing the average powder diameter can a
Contained 1 wt.% phosphate ester dispersant (Triton QS-44 from
increase the velocity of depositing droplets, which yields
Sigma-Aldrich; St. Louis, MO)
thinner lamellae (Ref 12). Thus, with the ability to spray

818—Volume 20(4) June 2011 Journal of Thermal Spray Technology


coatings will be referred to here as 2, 8, and 11 N coatings, 2.2 Thermal Conductivity Measurement

Peer Reviewed
respectively, where ‘‘N’’ stands for nanoparticle. After
Thermal diffusivity values were determined for 2, 8,
formulation the suspensions were milled for 3 h on a
and 11 N coatings. These measurements were conducted
rolling jar mill; the grinding media employed for this
using the laser flash system located at the High Temper-
process were 9.5-mm diameter by 9.5 mm-long cylinders
ature Materials Laboratory on the Oak Ridge National
of dense YSZ. Following milling no statistical difference
Laboratory (Oak Ridge, TN) campus. Test specimens
existed between the number-based d50 particle size for 2,
consisting of 12.7-mm diameter discs were cut into SPS
8, and 11 N suspensions.
coatings and removed from the substrate by dissolving the
Coatings were fabricated at Ames Laboratory (Ames,
copper in nitric acid. Before testing, these specimens were
IA) by spraying directly on 24-grit alumina-blasted copper
coated with graphite because of the YSZ transparency to
substrates that were 2.5-cm wide 9 10-cm long 9 0.5-cm
the laser radiation. The test specimens were oriented such
thick. Suspensions were agitated prior to and during the
that the effective direction of heat transfer through the
spray process using a combination of manual mixing,
microstructure was parallel to the coating spray direction.
ultrasonics, and a magnetically driven stir bar. Further-
The diffusivity values calculated from the temperature
more, all suspensions were filtered through a 90 lm sieve,
change were converted to conductivities using YSZ spe-
while being added to the pressure vessel used to supply the
cific heat data and the bulk density of each sample. Bulk
suspension to the plasma torch.
density and total porosity was measured via an Archime-
The experimental conditions briefly reviewed below
des approach. Additional details about this process have
were determined from a parametric study of the SPS
been published previously (Ref 28).
process employed for this work and were found to mini-
mize the volume fraction of unmelted YSZ powder
2.3 Small Angle Scattering Experiments
observed within the coatings (Ref 28, 29). The plasma
torch was a Praxair (Indianapolis, IN) SG-100 configured Small angle x-ray scattering (SAXS) experiments were
with a 3083-129 cathode, 3083-175 anode, 3083-113 gas performed on 2 and 8 N coating samples at the Advanced
injector, and a 6-mm diameter nozzle. The torch was Photon Source facility within Argonne National Labora-
operated at a current of 1000 A with primary and sec- tory (Argonne, IL) to characterize the void surfaces within
ondary plasma forming gas flows of 20 slpm Ar and these microstructures. Because of the high demand for the
60 slpm He, respectively, which produced a gun voltage of beamline needed to generate SAXS data, a second round
50 V. The spray direction describing the plasma flow of experiments could not be conducted to characterize
exiting the gun was perpendicular to the grit-blasted sur- 8 M and 11 N microstructures. The samples investigated
face. During deposition, the torch rastered horizontally at were obtained by using a robotic diamond saw to cut 1-
29 cm/s across this substrate surface; consecutive passes mm wide sections along the 12.7-mm diameter of the disc
were separated by 3 mm. The backside of the substrate specimens evaluated in the laser flash diffusivity testing.
was cooled using air jets with a combined volumetric flow Each section was epoxyed to a bolt so that it could be
rate of roughly 2700 lpm. A steel shroud surrounding the attached to an electromechanical positioning system and
substrate prevented the cooling air from interacting with placed into a 100-lm wide by 30-lm tall beam of colli-
the depositing YSZ droplets. The substrate temperature mated, monochromatic x-rays. In these experiments, the
was not monitored during spraying. beam was directed onto a coating sample along a path that
Table 2 lists the spray parameters for the SPS was perpendicular to the spray direction. Comparing the
experiments conducted during this study. All suspen- signal strength detected before and after the coating sec-
sions were injected as coherent streams orthogonal to tion permitted the determination of sample positions that
the plasma flow. These streams were generated by using resulted in the beam passing through the coating section.
nitrogen gas to force the suspension within a pressure By manipulating the sample position, SAXS data were
vessel through plastic tubing connected to a 230-lm recorded from a series of adjacent coating volume seg-
diameter ruby orifice. This orifice was purchased from ments; for each of the samples tested, the total coating
Bird Precision (Waltham, MA) and mounted on the volume evaluated was roughly 500 9 350 9 1000 lm.
internal injection port of the anode body. The minimum As the x-rays passed through a sample, differences in
final coating thickness desired in all experiments was the electron densities at the interfaces between the cera-
300 lm. mic and the gas within cracks and pores coherently scat-
tered waves along small angles from the beam direction
(Ref 30). Some of these waves did not undergo any further
scattering, and, as a result, the associated intensities were
correlated via a Porod constant to the total void specific
Table 2 Experimental spray parameters surface area (SSA) within the beam interaction volume
(Ref 5, 31). In addition, by recording the scattered inten-
Suspension 8M 2N 8N 11 N
sities with a two-dimensional detector, the void SSA var-
Standoff, cm 10 5 5 5 iation in directions perpendicular to the beam was plotted
Injection pressure, MPa 29 45 45 45 from the intensities of the singly scattered waves. The
Suspension flow, mL/min 42 52 52 52
Coating thickness, lm/pass 4 1 6 6
direction of each of these scattering events was perpen-
dicular to the void surface (Ref 31, 32). Therefore, the

Journal of Thermal Spray Technology Volume 20(4) June 2011—819


data in a given direction on the 2-D intensity plot was the
Peer Reviewed

result of voids aligned along an orthogonal direction.


For each 2 and 8 N sample, scattering data was col-
lected from only one sample orientation because plasma-
sprayed coatings demonstrate rotational symmetry along a
spray direction perpendicular to the substrate (Ref 31). As
a result, orienting each sample so that the incident beam
was normal to the spray direction yielded results which
could be rotated to form a complete description of scat-
tering from the entire the sample void system (Ref 31).
However, because they were not confirmed by conducting
scattering experiments at multiple sample orientations,
the term apparent is applied to the void SSA results for
the 2 and 8 N samples per the convention of Ilavsky et al.
(Ref 31). A further description of the setup and conditions
for these experiments has been published elsewhere
(Ref 28).

3. Results
3.1 Microstructural Description
Figure 1 shows that the 8 M coating microstructures
were the same as those of a typical APS coating, consisting
of overlapping lamellae that were ~10 lm and larger (i.e.,
longer than the image width) in diameter and several
microns thick. Lamellae often contained multiple intrala-
mellar cracks (Fig. 1a), and interlamellar pores were
observed between lamellae (Fig. 1b). Given the similarity
to APS microstructures, solvent evaporation during the
8 M experiments does not appear to have affected the
coating microstructure morphology.
As shown in Fig. 2, the SPS coatings produced from
nanopowder suspensions were also composed of lamellae.
Fig. 1 SEM images from an 8 M coating showing (a) a high
These lamellae were significantly smaller than those in the magnification of the top surface and (b) a fractured cross section
8 M coatings, having single digit micron diameters and normal to the substrate
thicknesses in the hundreds of nanometers. Among the
nanopowder feedstock coatings, the lamella diameters
appeared to increase with suspension powder concentra- of lamellae visible in micrographs of the 2 N (not shown),
tion. For example, the 11 N microstructure in Fig. 2(c) 8 N (Fig. 2a), and 11 N (Fig. 2c) top surfaces did not
exhibits lamellae with larger diameters than those shown contain cracks. In fact, intralamellar cracking was only
in the 8 N microstructure of Fig. 2(a). As the particle size observed among the largest YSZ lamellae within the 8 and
measurements showed the nanopowder suspensions to 11 N microstructures. The difference in the contraction
have equivalent d50 values, a difference in lamella size is along the lower and upper surfaces during cooling will
likely due to powder concentration differences in the increase with the lamella thickness. Consequently, in
fragmented suspension droplets within the plume. The comparison to the 8 M lamellae, smaller strains were
evaporation/combustion of the solvent in these droplets generated within the thinner diameter 2, 8, and 11 N
will eventually cause the powders to touch (Ref 13, 33). lamellae, which decreased intralamellar cracking in these
On complete melting, the resulting powder agglomerate microstructures.
will form one droplet. Agglomerates composed of an in- As shown in Fig. 3 and 4, the differences between
creased volume of powders, as would be the case for 11 N nanometer and micrometer feedstock fabricated coatings
suspensions compared to 8 N suspensions, should lead to went beyond lamella dimensions and cracking. Cauli-
larger droplets. Thus, the suspension powder concentra- flower-like cluster formations comprised the top surfaces
tion in SPS can influence the size of coating microstruc- of the 2 N (Fig. 3a), 8 N (Fig. 3b), and 11 N (Fig. 3c)
tural features in a manner analogous to the powder coatings, although the cluster formations are noticeably
feedstock diameter in APS (Ref 34). less distinct on the top surface of 11 N coatings than on
Lamellar pores still existed between lamellae in the 2, those of the 2 or 8 N coatings. Cross-sectional micro-
8, and 11 N coatings, an example of which can be seen in structures exhibited bands of porosity which resulted in
Fig. 2(b). However, unlike the 8 M coatings, the majority columnar YSZ structures with diameters of ~100 lm

820—Volume 20(4) June 2011 Journal of Thermal Spray Technology


Peer Reviewed
Fig. 2 SEM images showing (a) the top surfaces of 8 N lamellae, (b) a lamellar pore between 8 N lamellae, (c) the top surfaces of 11 N
lamellae, and (d) cross sections of fractured 8 N lamellae

(Fig. 4a). In the 2 and 8 N coatings, as shown in Fig. 4(b) investigations (Ref 5, 31, 32) have previously examined
and (c), the columns tended to extend through the coating the void SSA orientation in APS YSZ coatings produced
thickness. The cross-sectional microstructure of the 11 N with a spray direction orthogonal to the substrate surface.
coating was subtly different from those of the 2 and 8 N The reported findings should be applicable to the 8 M
coatings in that the 11 N porosity bands tended to merge coatings in the current research. In fact, the work by
(Fig. 4d). This change resulted in fewer columnar struc- Petorak (Ref 32) used the same powder as that employed
tures spanning the thickness of the 11 N coatings. to produce the 8 M coatings. The SANS data in these
studies show the APS YSZ coating samples contained
3.2 Small Angle Scattering Results interlamellar pores that were primarily oriented parallel to
the substrate and intralamellar cracks which were mostly
Figure 5 is the SAXS data for an as-sprayed (AS) 8 N perpendicular to the substrate.
coating. The near circular azimuthal dependence of the
Porod constants calculated from the singly scattered x-ray
3.3 Thermal Conductivity Results
intensities is representative of SAXS data obtained for the
AS 2 N coatings previously published (Ref 28), and 2 and As an initial evaluation of thermal barrier perfor-
8 N coatings after heat-treatment (not shown) (Ref 28, mance, the thermal conductivity (kth) of 2, 8, and 11 N
29). This result indicates that the combined apparent SSA coating samples were measured. Figure 6 shows temper-
of cracks and pores within these coatings was approxi- ature-dependent thermal conductivity results from repre-
mately isotropic. In addition, the SAXS results showed sentative SPS samples tested in the AS condition, and
that the total apparent void SSA decreased by ~30% after after being heat treated (HT) via a box furnace in air for
50 h at 1200 C, suggesting sintering occurred during the 50 h at 1200 C to simulate TBC service in a gas turbine.
heat treatment (Ref 29). For comparison, corresponding data replotted from Ref 7
SAXS experiments were not conducted on 8 M coating and 10 on APS and EB-PVD samples, respectively, is also
samples. However, small angle neutron scattering (SANS) included in Fig. 6.

Journal of Thermal Spray Technology Volume 20(4) June 2011—821


Peer Reviewed

Fig. 4 (a) Optical cross-sectional micrograph of a 2 N coating


showing columns. Cross-sectional SEM images showing the (b)
columnar structures in a fractured 2 N cross section, (c) porosity
Fig. 3 SEM images showing the top surfaces on (a) 2 N, (b) bands in a polished 8 N cross section, and (d) porosity bands in a
8 N, and (c) 11 N coatings polished 11 N cross section

These data show that, in the AS condition, ranking the 4. Discussion


coating types from lowest to highest kth gives APS, 11 N,
EB-PVD, 2 N, and 8 N, with the 11 N and EB-PVD types
4.1 Microstructural Development
being basically equivalent. All heat treatment kth data
exhibited an increase over the corresponding AS values. While the columnar and cauliflower 2, 8, and 11 N
The 2 N, 11 N, and EB-PVD values increased by microstructural features are not typical of spray-deposited
approximately 0.2 W/m/K, while the APS kth measure- coatings, some SPS studies (Ref 14, 20, 35) have reported
ments went up on average by 1 W/m/K. these structures. In addition, similar features have been

822—Volume 20(4) June 2011 Journal of Thermal Spray Technology


which the spray direction was ~39 from the substrate

Peer Reviewed
normal (Ref 36). The columns were separated from where
the spray plume contacted the plate and were attributed to
satellite droplets generated during the impact of the pri-
mary droplets on the substrate (Ref 36). Kanouff et al.
proposed that the spray gases flowing along the plate
surface carried the smaller satellite droplets downstream
of the main deposition region and deposited them on
surface asperities. As the deposition occurred preferen-
tially on the side of the asperity facing the main deposition
region, the deposits grew as columns into the gas flow
transporting the satellite droplets. Hass et al. used the
electron-beam directed vapor deposition (EB-DVD)
process to produce cauliflower-like top surfaces on YSZ
coatings (Ref 8). In these experiments, the angle between
the vapor jet and the substrate surface was oscillated be-
tween 135 and 135, causing most vapor to move along
directions off the substrate normal prior to condensation
(Ref 8). Findings like those published in Ref 8 and 36
suggest that the unique characteristics of the 2, 8, and 11 N
microstructures resulted from the deposition of YSZ
droplets having a significant velocity component parallel
Fig. 5 A plot showing the azimuthal dependence of the Porod to the substrate.
constants (units of cm 5) calculated from the SAXS of a beam For a spray deposition process, decreasing the particle,
perpendicular to the spray direction in an as-sprayed 8 N coating. and hence, the droplet size in the gas/plasma plume in-
The Porod shape describes the apparent distribution of the void
specific surface area in a plane perpendicular to the substrate. creases the sensitivity of these bodies to changes in gas/
The 0/180 axis in the above image is parallel to the substrate plasma velocity. In this study, the plasma flow was direc-
ted normal to the substrate. As a result, near the coating/
substrate the impinging plasma turned and moved across
this surface. This change in the plasma flow exerted drag
forces on the YSZ that tended to slow the droplet velocity
component perpendicular to the substrate and increase
the velocity component parallel to the substrate. Using
numerical modeling, Oberste Berghaus et al. calculated
the effects these forces have on various sizes of zirconia
particles initially traveling at 400 m/s perpendicular to the
substrate and within a plasma flow of equal speed and
direction (Ref 16). These particle velocities are consistent
with droplet velocities measured at 5 cm from the gun face
during SPS experimentation without a substrate (Ref 16).
The results of the model, which have been replotted in
Fig. 7, show that the inertia of a 40-lm diameter zirconia
particle traveling at 400 m/s is large enough to be unaf-
fected by drag from the impinging plasma; however, these
drag forces were determined to affect the trajectories of
400 m/s zirconia particles with diameters between 1 and
5 lm (Ref 16).
Using SEM micrographs like Fig. 2(b) and (d), the
average lamella thickness for all the nanopowder SPS
Fig. 6 Thermal conductivity vs. temperature data for as-sprayed
(open points) and heat-treated (filled points) samples from 2, 8, coatings was measured to be 300 ± 30 nm, where the
11 N, APS, and EBPVD coatings. The APS and SPS HT data are interval width captures the central 95% of a t-distribution
from samples heated for 50 h at 1200 C; EBPVD HT data are describing the lamella thicknesses. The average lamella
from a sample heated for 100 h at 1100 C. 2 N data are replotted diameter in 8 N coatings was estimated from a series of
from Ref 28; APS data are replotted from Ref 10, and EBPVD
data are replotted from Ref 7
micrographs like Fig. 2(a). An examination of this repre-
sentative figure shows that the longest dimension visible
for the majority of lamellae is <3.0 lm. Consequently,
observed previously in coatings generated by other spray 3.0 lm was used as the average lamella diameter to esti-
or spray-like processes. Kanouff et al. found isolated mate conservatively the size of the YSZ droplets depos-
columnar deposits on a grit-blasted aluminum plate sub- ited while spraying an 8 wt.% nanopowder suspension.
strate used in a high velocity oxy-fuel spray experiment for From the above dimensions, the lamellae were formed by

Journal of Thermal Spray Technology Volume 20(4) June 2011—823


Peer Reviewed

Fig. 7 Numerical simulation results replotted from Ref 16


showing how the changes in plasma velocity stemming from
substrate impingement and the resulting drag influence the sub-
strate normal velocity component of different diameters of zir-
conia particles within the plasma
Fig. 8 Schematics illustrating the deposition characteristics
occurring on and away from substrate surface asperities during
YSZ droplets with diameters <2 lm. Therefore, com- (a) 1SD, (b) 2SD, and (c) 3SD. The spray direction is the same in
paring the droplet size to the results in Fig. 7 suggests that each figure. The black regions denote the substrate. The light
drag forces from the impinging plasma caused the flight gray regions represent material deposited by droplets having a
substrate parallel velocity component directed from left to right,
paths of all the YSZ droplets in the 8 N experiments to and the dark gray regions represent material deposited by
deviate from the spray direction. Conversely, the plasma droplets having a substrate parallel velocity component directed
impingement should have produced little change in the from right to left
paths of droplets resulting from the 15-lm diameter
powder.
The impact trajectory predictions above led to the idea Consequently, 1SD results almost exclusively in droplet
that the 2, 8, and 11 N microstructural features, which are impacts on the sides of surface asperities, as shown in
unique from common spray-deposited coatings, result Fig. 8(a). Initially, deposits formed on substrate asperities
from directional changes in droplet travel near the sub- grow both laterally and vertically. During this time, the
strate. A similar hypothesis was also suggested by other lateral growth of taller deposits may engulf shorter depos-
researchers that found substrate roughness and size its. Among the remaining deposits of similar heights, the
influenced the microstructures of coatings produced dur- increase in drag forces generated by the decrease in spacing
ing SPS with nanosized powders (Ref 35). To develop this separating these structures reduces the plasma flowing be-
premise further, theories were formulated defining depo- tween them. This process eventually causes the lateral
sition mechanisms leading to the three different micro- growth to stop, leaving an interdeposit gap. As spraying
structure types, i.e., those apparent in the 2 N/8 N, 11 N, proceeds, the ongoing vertical growth of the deposits and
and 8 M coatings, observed in this study. For these theo- interdeposit gaps produces columnar structures that span
ries, the YSZ droplet velocity is 400 m/s perpendicular to the coating cross section normal to the substrate.
the substrate surface prior to any impingement effects, Spraying will naturally yield localized nonuniform dis-
and the substrate asperity heights are assumed to be sig- tributions of droplet impacts, which can lead to asperities
nificantly larger than the impacting droplet diameters. The in the coating surface. These coating asperities can be-
hypothetical deposition mechanisms outlined below may come sites of preferential 1SD. Consequently, the coating
change for conditions other than those assumed here. is expected to have a surface featuring multiple deposits
4.1.1 Type 1 Spray Deposition Microstructure Devel- separated by valleys; these deposits are the cluster for-
opment. Microstructures like those of the 2 and 8 N coat- mations comprising the cauliflower-like surface structures
ings (as revealed in Fig. 3a and b, and Fig. 4a-c) are mentioned previously.
proposed to form by a mechanism which will be referred to 4.1.2 Type 2 Spray Deposition Microstructure Devel-
as type 1 spray deposition (1SD). In 1SD, YSZ droplets opment. Type 2 spray deposition (2SD) is proposed as
move with the plasma through the initial turn at substrate describing conditions in which the droplet retains a
impingement. As a result, the substrate parallel velocity velocity directed primarily perpendicular to the substrate,
component of these droplets dominates the substrate nor- yet the substrate parallel component still affects deposi-
mal component. However, as the plasma flows across the tion. In this case, the droplets are more massive than those
substrate, the inertia difference between the plasma and depositing according to the 1SD mechanism, and they
small YSZ droplets cause some of the entrained droplets to separate from the plasma before completing the substrate
be unable to follow the more sudden directional changes impingement turn. Consequently, deposition in regions
produced as the plasma moves around surface asperities. between surface asperities increases with 2SD. However,

824—Volume 20(4) June 2011 Journal of Thermal Spray Technology


the effects of the impinging plasma drag cause the droplets 4.2 SAXS Analysis with Respect to Deposition

Peer Reviewed
to follow trajectories that result in surface asperities Conditions
shadowing portions of the substrate downstream from the
As shown in Fig. 1(b), an interlamellar pore has an
asperity, which is illustrated in Fig. 8(b). Early in the
orientation similar to the YSZ lamella formed by the
coating process, the shadowing occurring with 2SD will
droplet which created the pore. Therefore, knowledge of
create variations in the growth rate. Figure 8(b) shows the
the orientation of the interlamellar pores obtained from
thickness of the coating deposited in the unshadowed re-
small angle scattering experiments can be used to
gions at the left and right extremes of the figure to be
approximate the lamellae orientations within a coating.
growing twice as fast as that on the substrate asperities.
These experimental results will also include scattering
This difference is because the left substrate asperity blocks
from intralamellar cracks, but, since these cracks tend to
the oblique droplet trajectories from reaching two
run perpendicular to interlamellar pores, the scattering
regions—identified by an absence of light gray bands on
from these two void types will be orthogonal. In addition,
Fig. 8(b)—of the substrate to the right of the peak on this
based on SEM observations, intralamellar cracks are
asperity. Likewise, the right substrate asperity blocks
likely to constitute less of the total void surface area
droplets from reaching two regions—identified by an ab-
than interlamellar pores in nanopowder SPS coatings.
sence of dark gray bands on Fig. 8(b)—of the substrate to
Figure 1(b) shows that plasma-sprayed microstructures
the left of the peak on this asperity. As the height differ-
also contain spherical, intralamellar pores. However, the
ential between the coating in the unshadowed regions and
contribution from this type of pore can be neglected be-
the material deposited on top of the substrate asperities
cause their surface area is small relative to that of intral-
decreases, the deposition on the side of the former will
amellar cracks and interlamellar pores (Ref 5).
increase. This increase causes the unshadowed sections of
The azimuthal dependence of the Porod constants
the coating to grow toward that covering the substrate
calculated using SAXS intensities produced by 2 and 8 N
asperities. Therefore, given a long enough spray time, the
samples were near circular shapes (Fig. 5) (Ref 28).
coating building from the initially unshadowed sections
However, SEM images like Fig. 2(b) confirmed that the
will overgrow that above the substrate asperities. Due to
interlamellar pores in the 2 and 8 N coatings had oblate or
this overgrowth behavior, a coating microstructure formed
lamellar shapes. Therefore, the most probable explanation
by 2SD should exhibit a convergence of the porosity bands
for the isotropy of apparent void SSA in these SPS coat-
separating columnar structures like that observed in the
ings is a near random orientation of the lamellae relative
11 N coatings and shown in Fig. 4(d).
to the substrate plane. Such an orientation is expected for
As with 1SD, the droplet trajectories in the 2SD
coatings generated by 1SD because the droplet impacts in
mechanism promote cluster formations on coating surface
this deposition mechanism result from the plasma flowing
asperities. However, these structures exhibit less growth
around and over surface asperities. Consequently, the
with 2SD than with 1SD because droplets with velocities
planes describing droplet spreading should form a range of
mainly normal to the surface are less likely to deposit
angles with the substrate plane.
preferentially on surface asperities. Therefore, cluster
SANS results correlated to the 8 M coatings suggest
formations will be less distinct on the surface of a micro-
these coatings were composed primarily of lamellae with
structure produced by 2SD than by 1SD.
radial axes approximately parallel to the spray direction.
4.1.3 Type 3 Spray Deposition Microstructure Devel-
Such a microstructure is consistent with 3SD. This mech-
opment. Type 3 spray deposition (3SD) is proposed as
anism tends to generate planar coatings, which increases
describing conditions in which plasma drag forces do not
the probability of an impacting droplet spreading parallel
affect the deposition characteristics, meaning the droplets
to the substrate plane.
yielding 3SD are more massive than those causing 2SD. As
SAXS experiments were not performed on 11 N sam-
a result, 3SD is characterized by deposition on the entire
ples. Based on the earlier deposition theory, the void SSA
substrate surface (Fig. 8c) during each plasma gun pass, i.e.,
orientation in the 11 N coatings could be more anisotropic
droplets do not preferentially impact on surface asperities,
than those of the 2 or 8 N samples, with a greater SSA
and no shadowing occurs from these asperities. The droplet
percentage orientated roughly parallel to the substrate
trajectories required for this complete coverage would be
plane. However, the cauliflower-like surface features and
approximately normal to the substrate surface. Thus, Fig. 7
the kth increase with heat treatment (discussed below)
suggests that 3SD would occur when plasma spraying with
suggest the azimuthal dependence of Porod constants
YSZ droplets having diameters larger than 5 lm.
determined from the SAXS generated by 11 N micro-
The complete substrate coverage possible with 3SD
structures would be nearly isotropic.
prevents the formation of columnar structures separated
by porosity bands (Fig. 8c). In addition, the momentum of
4.3 Thermal Conductivity Discussion with Regard
the droplets in this mechanism should preferentially drive
to 1SD, 2SD, and 3SD Microstructures
liquid into the valleys in the coating/substrate surface.
Thus, coating asperities generated by nonuniform 3SD As the maximum temperature during the laser flash
will not grow into cluster formations; instead, the coating testing was 1200 C, the heat transfer occurred primarily
surface will tend to become more planar than the grit- by phonon transport in the YSZ (Ref 22). Voids within the
blasted substrate surface. This type of deposition produces microstructure inhibit this transport in two ways. First,
the stereotypical APS coating microstructure. because phonons must move around voids, these features

Journal of Thermal Spray Technology Volume 20(4) June 2011—825


Table 3 Bulk density and total porosity values APS YSZ samples can be attributed to the influence of
Peer Reviewed

for the SPS thermal conductivity test specimens void structure on this property. In the as-deposited con-
dition, the columnar grains forming an EB-PVD YSZ
Coating ID Bulk density, g/cm3 Total porosity, vol.% coating contain rows of isolated pores which are submi-
2N 3.92 ± 0.06 37 ± 2 cron in size and separation (Ref 7). Sintering of these
2 N-HT 4.20 ± 0.08 31 ± 2 coatings causes pore spheriodization, during which some
8N 5.19 ± 0.06 14 ± 1 pores grow at the expense of others (Ref 7). However,
11 N 4.50 ± 0.04 26 ± 1
11 N-HT 4.56 ± 0.03 25 ± 1
even after 100 h at 1100 C, the EB-PVD YSZ void
structure still consists of isolated pores that are submicron
Each uncertainty interval captures the central 95% of a t-distribution
in size and separation (Ref 7). On the other hand, the
describing the sample data
intralamellar cracks in an AS APS YSZ coating can link
interlamellar pores to produce void networks (Ref 31,
38). The destruction of these networks via crack closure
increase the average transport distance, and, second, by during sintering can leave isolated interlamellar pores
constricting the cross-sectional area available for trans- separated by millimeter-sized gaps, significantly raising
port, voids increase phonon-phonon scattering interac- the coating kth (Ref 32, 38). In fact, the results of Ref 10
tions (Ref 23). As stated in Section 2.2, the macroscopic and 32 show that roughly half of the kth increase for APS
heat flux through the test samples was one-dimensional YSZ coatings after a 3-50 h, 1200 C heat treatment can
and parallel to the coating spray direction. Therefore, both be attributed to crack closure. As for the relative kth in-
the amount and distribution of porosity within a micro- creases associated with pore spheroidization, APS YSZ
structure affected the sample thermal conductivity interlamellar porosity tends be oriented more normal to
(Ref 37). the spray direction than the intracolumnar porosity is to
A comparison of the density and porosity values in the EB-PVD column growth direction (Ref 6, 7, 31, 32,
Table 3 and the discussion of porosity orientation in the 38). Consequently, when the macroscopic heat flow
preceding section lead to the conclusion that the kth data direction is parallel to the APS coating spray direction
of the 2 N coatings was lower than that of the 8 N coatings and the EB-PVD column growth direction, pore spher-
because the former contained more porosity. Conversely, iodization will have a greater effect on the kth of the
the 11 N sample contained less porosity than the 2 N former coating type.
sample, yet the 11 N porosity distribution caused this The void structure of the nanopowder SPS coatings in
specimen to have lower kth values. The 11 N microstruc- this work was more typical of an EB-PVD YSZ coating
tures were the only ones observed to contain porosity than an APS YSZ coating. The extent of intralamellar
bands that interconnect, which creates more conduction cracking present in micrographs of these SPS coatings
pathway restrictions than the linear 2 N/8 N bands. Since suggests crack-pore void networks were rare, meaning the
the porosity bands are significantly larger than intrala- interlamellar pores were mainly isolated like intracolum-
mellar cracks or interlamellar pores, it is not surprising nar EB-PVD pores. In addition, the SAXS results indicate
that the 11 N porosity was more effective at reducing heat the majority of intralamellar pore SSA to be aligned along
conduction than that of the 2 N specimen. directions that cannot be considered roughly perpendicu-
The increase in coating thermal conductivity data lar to the spray direction. As a result of the similarities, the
following heat treatments at 1100 and 1200 C suggest kth increase seen in Fig. 6 between the AS and HT SPS
microstructural void changes occurred during heating. At samples shows a better agreement with the corresponding
these temperatures a YSZ microstructure will undergo change in the EB-PVD data than that for the APS data.
sintering, resulting in crack closure and pore spherodi-
zation (Ref 5, 10, 38, 39). These processes resulted in a
30% decrease in the total apparent void SSA observed
between the AS and heat treated 2 and 8 N SAXS 5. Summary
samples (Ref 29). Similarly, Petorak (Ref 32) reports the
total apparent void SSA in APS YSZ coatings decreased The SPS process was employed to produce coatings
by 16% after 3 h at 1200 C. Following 100 h at 1100 C, from suspensions containing 15 lm and 80-nm diameter
the SSA of the open and closed porosity in the EB-PVD YSZ powders. Suspensions containing 8 wt.% of 15 lm
YSZ coatings of (Ref 40) each decreased by 68%. Al- powder (8 M) resulted in the well-known spray deposition
though each coating type exhibited a void SSA reduc- microstructure—overlapping lamellae primarily oriented
tion, the corresponding average kth increases do not parallel to the substrate. Conversely, the 80 nm powder
correlate linearly with the SSA changes. As shown in suspensions yielded unique microstructures in which
Fig. 6, the kth values for the 2 N and EB-PVD samples lamellae formed columnar structures topped with cauli-
increase an average of 0.2 W/m/K between the AS and flower-like formations. When spraying suspensions con-
HT measurements, while Ref 32 reports a 0.54 W/m/K taining 2 wt.% (2 N) or 8 wt.% (8 N) nanopowder, the
average kth increase after heating APS YSZ samples for columns were composed of an approximately random
3 h at 1200 C. orientation of lamellae and were separated by linear
The disparity in the sensitivity of coating thermal porosity bands. With an 11 wt.% (11 N) powder suspen-
conductivity to heat treatment between EB-PVD and sion, interconnected porosity bands were observed.

826—Volume 20(4) June 2011 Journal of Thermal Spray Technology


Hypotheses on the formation mechanisms producing Laboratory User Program, Oak Ridge National Labora-

Peer Reviewed
these coatings were developed from microstructural tory, managed by UT-Battelle, LLC, for the U.S.
observations and predictions of droplet flight paths. This Department of Energy under contract number DE-AC05-
information suggests that the change in relative influences 00OR22725.
of drag and inertial forces with YSZ droplet size can result
in three types of spray deposition. For 1SD, the plasma
drag forces during substrate impingement dominate the
References
droplet inertia and redirect the droplet velocity from
normal to along to the substrate surface. Consequently, 1. R.A. Miller, Thermal Barrier Coatings for Aircraft Engines:
droplets impact preferentially on asperities, generating History and Directions, J. Therm. Spray Technol., 1997, 6(1),
deposits that grow to become columnar structures sepa- p 35-42
2. U. Schulz, C. Leyens, K. Fritscher, M. Peters, B. Saruhan-Brings,
rated by linear porosity bands. The 1SD microstructure is O. Lavigne, J.-M. Dorvaux, M. Poulain, R. Mévrel, and M.
analogous to that of the 2 and 8 N coatings from this Caliez, Some Recent Trends in Research and Technology of
study. With 2SD, the droplet velocity remains mostly Advanced Thermal Barrier Coatings, Aerosp. Sci. Technol., 2003,
normal to the substrate surface, but the impinging plasma 7, p 73-80
3. P. Fauchais, Understanding Plasma Spraying, J. Phys. D Appl.
drag influences the droplet trajectories such that surface Phys., 2004, 37, p R86-R108
asperities block deposition in downstream regions and 4. R. McPherson, A Review of Microstructure and Properties of
create porosity bands. The coating growth rate differences Plasma Sprayed Ceramic Coatings, Surf. Coat. Technol., 1989,
between the resulting shadowed and unshadowed regions 39(40), p 173-181
cause faster growing deposits to overgrow slower ones, 5. J. Ilavsky, G.G. Long, A.J. Allen, and C.C. Berndt, Evolution of
the Void Structure in Plasma-Sprayed YSZ Deposits During
producing a convergence of porosity bands like what was Heating, Mater. Sci. Eng. A, 1999, 272, p 215-221
observed in the 11 N coatings. In 3SD, the coating 6. A. Kulkarni, A. Goland, H. Hermana, A.J. Allen, T. Dobbins,
microstructure exhibits no signs of droplets being influ- F. DeCarlo, J. Ilavsky, G.G. Long, S. Fang, and P. Lawton,
enced by drag from the plasma impingement. These Advanced Neutron and X-ray Techniques for Insights into the
Microstructure of EB-PVD Thermal Barrier Coatings, Mater. Sci.
droplets follow a path with a substrate normal component Eng. A, 2006, 426, p 43-52
larger enough to prevent any preferential deposition on 7. F. Renteria, B. Saruhan, U. Schulz, H.-J. Raetzer-Scheibe, J.
surface asperities or asperity shadowing. Thus, like the Haug, and A. Wiedenmann, Effect of Morphology on Thermal
8 M coatings, 3SD microstructures do not contain Conductivity of EB-PVD PYSZ TBCs, Surf. Coat. Technol.,
columnar structures or porosity bands, and the lamellae 2006, 201, p 2611-2620
8. D.D. Hass, A.J. Slifka, and H.N.G. Wadley, Low Thermal Con-
tend to be oriented parallel to the substrate. ductivity Vapor Deposited Zirconia Microstructures, Acta Ma-
The thermal diffusivities of 80 nm powder SPS coatings ter., 2001, 49, p 973-983
were measured in as-sprayed and heat-treated conditions, 9. H. Zhao, F. Yu, T.D. Bennett, and H.N.G. Wadley, Morphology
and the data were used to calculate coating thermal con- and Thermal Conductivity of Yttria-Stabilized Zirconia Coatings,
Acta Mater., 2006, 54, p 5195-5207
ductivities. Over the temperature range from 700 to 10. R.W. Trice, Y.J. Su, J.R. Mawdsley, K.T. Faber, A.R. De
1200 C, the average thermal conductivities of 2 and 11 N Arellano-Lopez, H. Wang, and W.D. Porter, Effect of Heat
samples were, respectively, 1.4 and 1.2 W/m/K in the AS Treatment on Phase Stability, Microstructure, and Thermal
condition and 1.6 and 1.4 W/m/K after 50 h at 1200 C. Conductivity of Plasma-Sprayed YSZ, J. Mater. Sci., 2002, 37,
The kth increases for these SPS samples were less than p 2359-2365
11. P. Fauchais, M. Fukumoto, A. Vardelle, and M. Vardelle,
one-third of those for APS YSZ coatings tested under the Knowledge Concerning Splat Formation: An Invited Review,
same conditions (Ref 10, 32). The decreased change in the J. Therm. Spray Technol., 2004, 13, p 337-360
SPS samples kth after heat treatment was attributed to 12. A. Kulkarni, A. Vaidya, A. Goland, S. Sampath, and H. Herman,
randomly oriented and isolated interlamellar pores, Processing Effects on Porosity-Property Correlations in Plasma
Sprayed Yttria-Stabilized Zirconia Coatings, Mater. Eng. A, 2003,
implying the 2 and 11 N coatings had void structure 359, p 100-111
characteristics similar to an EB-PVD coating. 13. Z. Chen, R.W. Trice, M. Besser, X. Yang, and D. Sordelet, Air-
Plasma Spraying Colloidal Solutions of Nanosized Ceramic
Powders, J. Mater. Sci., 2004, 39, p 4171-4178
14. R. Siegert, ‘‘A Novel Process for the Liquid Feedstock Plasma
Acknowledgments Spray of Ceramic Coatings with Nanostructural Features,’’
Doctoral Thesis, Institut fur Werkstoffe und Verfahren der
Major portions of this research were funded by the Energietechnik, 2005
National Science Foundation via grant CMMI-0456534. 15. J. Fazilleau, C. Delbos, V. Rat, J.F. Coudert, P. Fauchais, and B.
Ames Laboratory is operated for the U.S. Department of Pateyron, Phenomena Involved in Suspension Plasma Spraying
Energy by Iowa State University under Contract No. DE- Part 1: Suspension Injection and Behavior, Plasma Chem. Plasma
AC02-07CH11358. Use of the Advanced Photon Source at Process., 2006, 26, p 371-391
16. J. Oberste Berghaus, S. Bouaricha, J.-G. Legoux, and C. Moreau,
Argonne National Laboratory was supported by the U.S. Injection Conditions and In-Flight Particle States in Suspension
Department of Energy, Office of Science, Office of Basic Plasma Spraying of Alumina and Zirconia Nano-Ceramics,
Energy Sciences, under Contract No. DE-AC02- Proceedings of the International Thermal Spray Conference,
06CH11357. This project involved research sponsored by C. Berndt and E. Lugsheider, Ed., May 2-4, 2005 (Basel,
Switzerland), ASM International, 2005
the Assistant Secretary for Energy Efficiency and 17. C. Delbos, J. Fazilleau, V. Rat, J.F. Coudert, P. Fauchais, and L.
Renewable Energy, Office of FreedomCAR and Vehicle Bianchi, Influence of Powder Size Distribution and Heat Flux on
Technologies, as part of the High Temperature Materials Yttria Stabilised Coatings Elaborated by Liquid Suspension

Journal of Thermal Spray Technology Volume 20(4) June 2011—827


Injection in a DC Plasma Jet, Proceedings of the International 29. K. VanEvery, ‘‘Development and Evaluation of Suspension
Peer Reviewed

Thermal Spray Conference, C. Berndt and E. Lugsheider, Ed., Plasma Sprayed Yttria Stabilized Zirconia Coatings as Thermal
May 2-4, 2005 (Basel, Switzerland), ASM International, 2005 Barriers,’’ Doctoral Thesis, Purdue University, 2009
18. J. Oberste Berghaus, S. Bouaricha, J.-G. Legoux, C. Moreau, and 30. O. Glatter and O. Kratky, Small Angle X-ray Scattering, Aca-
T. Chraska, Suspension Plasma Spraying of Nano-Cermics Using demic Press, New York, 1982
an Axial Injection Torch, Proceedings of the International Ther- 31. J. Ilavsky, A.J. Allen, G.G. Long, S. Krueger, C.C. Berndt, and
mal Spray Conference, C. Berndt and E. Lugsheider, Ed., May 2- H. Herman, Influence of Spray Angle on the Pore and Crack
4, 2005 (Basel, Switzerland), ASM International, 2005 Microstructure of Plasma-Sprayed Deposits, J. Am. Ceram. Soc.,
19. R. Siegert, J.-E. Doring, J.-L. Marques, R. Vassen, D. Sebold, 1997, 80(3), p 733-742
and D. Stover, Influence of the Injection Parameters on the 32. C.A. Petorak, ‘‘Evolution of the Plasma-Sprayed Microstructure
Suspension Plasma Spraying Coating Properties, Proceedings of in 7 wt% Yttria-Stabilized Zirconia Thermal Barrier Coatings
the International Thermal Spray Conference, C. Berndt and E. During Uniaxial Stress Relaxation and the Concomitant Changes
Lugsheider, Ed., May 2-4, 2005 (Basel, Switzerland), ASM in Material Properties,’’ Doctoral Thesis, Purdue University,
International, 2005 2007
20. C. Delbos, J. Fazilleau, V. Rat, J.F. Coudert, P. Fauchais, and B. 33. L. Pawlowski, Suspension and Solution Thermal Spray Coatings,
Pateyron, Phenomena Involved in Suspension Plasma Spraying, Surf. Coat. Technol., 2009, 203, p 2807-2829
Part 2: Zirconia Particle Treatment and Coating Formation, 34. R.W. Trice, C. Batson, C. Scharff, and K.T. Faber, The Role of
Plasma Chem. Plasma Process., 2006, 26, p 393-414 Starting Powder Size on the Compressive Response of Stand-
21. W. Beele, G. Marijnissen, and A. van Lieshout, The Evolution of Alone Plasma-Sprayed Alumina Coatings, J. Mater. Sci., 2002, 37,
Thermal Barrier Coatings—Status and Upcoming Solutions for p 629-636
TodayÕs Key Issues, Surf. Coat. Technol., 1999, 120-121, p 61-67 35. J. Oberste Berghaus, J.-G. Legoux, C. Moreau, F. Tarasi, and T.
22. J.R. Nicholls, K.J. Lawson, A. Johnstone, and D.S. Rickerby, Chráska, Mechanical and Thermal Transport Properties of Sus-
Methods to Reduce the Thermal Conductivity of EB-PVD TBCs, pension Thermal-Sprayed Alumina-Zirconia Composite Coat-
Surf. Coat. Technol., 2002, 151-152, p 383-391 ings, J. Therm. Spray Technol., 2008, 17(1), p 91-104
23. P.G. Klemens, Theory of Thermal Conduction in Thin Ceramic 36. M.P. Kanouff, R.A. Neiser, Jr., and T.J. Roemer, Surface
Films, Int. J. Thermophys., 2001, 22(1), p 265-275 Roughness of Thermal Spray Coatings Made with Off-Normal
24. Coulter LS Series Product Manual, Coulter Corporation, Mia- Spray Angles, J. Therm. Spray Technol., 1998, 7(2), p 219-228
mi, FL 1994 37. A. Bjorneklett, L. Haukeland, J. Wigren, and H. Kristiansen,
25. A. Jillavenkatesa, S.J. Dapkunas, and L-S. H. Lum, NIST Special Effective Medium Theory and the Thermal Conductivity of
Publication 960-1: Particle Size Characterization, http:// Plasma-Sprayed Ceramic Coatings, J. Mater. Sci., 1994, 29,
www.msel.nist.gov/practiceguides/SP960_1.pdf, 2001 p 4043-4050
26. S. Raghavan, H. Wang, R.B. Dinwiddie, W.D. Porter, and M.J. 38. A.J. Allen, J. Ilavsky, G.G. Long, J.S. Wallace, C.C. Berndt, and
Mayo, The Effect of Grain Size, Porosity and Yttria Content on H. Herman, Microstructural Characterization of Yttria-Stabilized
the Thermal Conductivity of Nanocrystalline Zirconia, Scr. Zirconia Plasma-Sprayed Deposits Using Multiple Small-Angle
Mater., 1998, 39(8), p 1119-1125 Neutron Scattering, Acta Mater., 2001, 49, p 1661-1675
27. J.R. Piascik, J.Y. Thompson, C.A. Bower, and B.R. Stoner, Eval- 39. K.A. Erk, C. Deschaseaux, and R.W. Trice, Grain-Boundary
uation of Crystallinity and Film Stress in Yttria-Stabilized Zirconia Grooving of Plasma-Sprayed Yttria-Stabilized Zirconia Ther-
Thin Films, J. Vac. Sci. Technol., 2005, A23(5), p 1419-1424 mal Barrier Coatings, J. Am. Ceram. Soc., 2006, 89(5), p 1673-
28. K. VanEvery, M. Krane, R.W. Trice, W. Porter, H. Wang, M. 1678
Besser, D. Sordelet, J. Ilavsky, and J. Almer, In-Flight Alloying 40. A. Flores Renteria and B. Saruhan, Effect of Ageing on Micro-
of Nanocrystalline Yttria-Stabilized Zirconia Using Suspension structure Changes in EB-PVD Manufactured Standard PYSZ
Spray to Produce Ultra-Low Thermal Conductivity Thermal Top Coat of Thermal Barrier Coatings, J. Eur. Ceram. Soc., 2006,
Barriers, Int. J. Appl. Ceram. Technol. (in press) 26, p 2249-2255

828—Volume 20(4) June 2011 Journal of Thermal Spray Technology

You might also like